Mercury(II) Complexes of the Carba-closo

Mercury(II) Complexes of the Carba-closo...
0 downloads 0 Views 4MB Size
Article pubs.acs.org/Organometallics

Mercury(II) Complexes of the Carba-closo-dodecaboranyl Ligands [closo-1-CB11X11]2− (X = H, F, Cl, Br, I) Alexander Himmelspach,† Jan A. P. Sprenger,† Jonas Warneke,*,‡ Manfred Zaḧ res,§ and Maik Finze*,† †

Institut für Anorganische Chemie, Julius-Maximilians-Universität Würzburg, Am Hubland, 97074 Würzburg, Germany Institut für Organische Chemie, Fachbereich 2 (Chemie/Biologie), Universität Bremen, Leobener Strasse, NW2, 28359 Bremen, Germany § Institut für Physikalische Chemie, Universität Duisburg-Essen, Universitätsstrasse 5, 45141 Essen, Germany ‡

S Supporting Information *

ABSTRACT: Salts of the mercury(II) complexes [Hg(closo-1CB11X11)2]2− (X = H (1), Cl (3), Br (4)) and [PhHg(closo-1CB11X11)]− (X = H (6), Cl (8), Br (9), I (10)) were synthesized and characterized by multi-NMR spectroscopy, mass spectrometry, elemental analysis, and differential scanning calorimetry. Single crystals of Cs21·2Et2O, Cs23·MeCN, Cs24·4Me2CO, Cs9, and [Et4N]6·0.5Me2CO were studied by X-ray diffraction, and the interpretation of the bond properties is supported by theoretical data. In contrast to the mercury atom of the previously published [Hg(closo-1-CB11F11)2]2− (2), which coordinates either acetonitrile or water, the metal atom of the related dianionic complexes 1, 3, and 4 does not reveal any further coordination. According to results derived from DFT and ab initio calculations, this different behavior is reasoned in the case of 1 by a reduced Lewis acidity at mercury and in the case of 3 and 4 by the increased shielding of the central mercury atom as a result of the bulky halogenated carba-closo-dodecaboranyl ligands [closo-1-CB11X11]2− (X = Cl, Br). The dianionic complex [Hg(closo-1-CB11I11)2]2− (5) with the bulkiest carba-closo-dodecaboranyl ligand was generated via collision-induced dissociation and characterized by (−)-ESI mass spectrometry. The fragmentation pathways of the anionic complexes [Hg(closo-1-CB11X11)2]2− (X = H, F, Cl, Br, I (1−5)) and [PhHg(closo-1-CB11X11)]− (X = H, F, Cl, Br, I (6−10)) were studied by (−)-ESI mass spectrometry.



CB11H11)]−.10 In the case of the homoleptic complex 2, the coordination of acetonitrile or water to the Lewis acidic mercury center was observed, and [Et4N]2[Hg(closo-1CB11F11)2(NCMe)] ([Et4N]2[2·NCMe]) and [Et4N]4[{Hg(closo-1-CB11F11)2}2(OH2)] ([Et4N]4[22·OH2]) were characterized by vibrational spectroscopy and single-crystal X-ray diffraction. Obviously, the fluorinated [closo-1-CB11F11]2− ligand results in a similar behavior of HgII in 2 as found for mercury complexes with perfluoroaryl and dicarba-closododecaboranyl ligands. This leads to the question of whether related mercury(II) complexes with carba-closo-dodecaboranyl ligands especially with perhalogenated clusters reveal an analogous behavior, or not. We are interested in this topic in connection with our study on the specific characteristics of highly fluorinated {closo-1-CB11} clusters in comparison to other halogenated and nonhalogenated {closo-1-CB11} derivatives.12−14 Here, we report on the synthesis of the cesium and [Et4N]+ salts of the HgII complexes [Hg(closo-1-CB11X11)2]2− (X = H

INTRODUCTION Transition-metal complexes of dicarba-closo-dodecaboranyl ligands with an exo-polyhedral metal−Ccluster σ bond are wellknown,1,2 and they are used in a number of applications, for example, in catalysis. Important examples are mercury(II) complexes with dicarba-closo-dodecaboranyl ligands,2−4 which have been investigated with all three isomeric clusters {closo1,2-C2B10}, {closo-1,7-C2B10}, and {closo-1,12-C2B10}, because they exhibit strong similarities to HgII complexes with electronpoor perfluoroaryl ligands.5,6 Both contain Lewis acidic mercury atoms that can coordinate further neutral as well as anionic ligands and are, therefore, discussed as anion receptors, materials for gas storage, catalysis, or as buildings blocks in supramolecular chemistry.2,4−6 To the best of our knowledge, only a few related complexes with carba-closo-dodecaboranyl ligands and a metal−Ccluster σ bond have been described, so far:7 [ClCu(closo-1-CB11F11)]2−,8 [Ag(closo-1-CB11I5Br6)2]3−,9 and [PhAu(closo-1-CB11H11)]−.10 Recently, we have reported on the tetraethylammonium salts of the mercury(II) complexes [Hg(closo-1-CB11F11)2]2− (2) and [PhHg(closo-1-CB11F11)]− (7) that contain a perfluorinated carba-closo-dodecaboranyl ligand.11 Both complexes exhibit high thermal and chemical stability, which is similar to the behavior of the anionic complex [Ph 3 PAu(closo-1© 2011 American Chemical Society

Special Issue: Fluorine in Organometallic Chemistry Received: October 23, 2011 Published: December 16, 2011 1566

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

Scheme 1. Synthesis of 1−4 and 6−10 with nBuLi as Base

(1), Cl (3), Br (4)) and [PhHg(closo-1-CB11X11)]− (X = H (6), Cl (8), Br (9), I (10)), which were characterized by multiNMR spectroscopy and elemental analysis. The crystal structures of the cesium salts of all three dianions and of anion 9 as well as of [Et4N]6 were studied by X-ray diffraction, and the discussion of the bond properties is supported by results derived from DFT and ab initio calculations. A possible explanation for the coordination of a further ligand solely to HgII of the fluorinated derivative 2 and not to its homologues 1, 3, and 4 is provided on the basis of theoretical data. All anions, including the fluorinated complexes [PhHg(closo-1-CB11F11)]− (7) and [Hg(closo-1-CB11F11)2]2− (2) as well as the iodinated derivative [Hg(closo-1-CB11I11)2]2− (5), which was generated in an ion trap, were studied by (−)-ESI mass spectrometry, and the fragmentation patterns of the anions obtained from these HgII complexes are discussed in detail.

Scheme 2. In Situ Synthesis of 3 and 4

The synthesis of salts of the dianion [Hg(closo-1-CB11I11)2]2− (5) using either route outlined in Schemes 1 and 2 was not successful, so far. The NMR spectroscopic analysis of the reaction mixtures provided some evidence for the formation of products that contain Hg−Ccluster σ bonds. No cleavage of B−I bonds was observed. These results were confirmed by (−)-ESI as well as (−)-MALDI mass spectrometric studies, and finally the formation of dianion 5 was achieved in an ion trap, as outlined in the following section that describes the mass spectrometric experiments in detail. The decomposition of the Cs+ salts of 1−4 starts above 200 °C, as determined by differential scanning calorimetry (DSC). The 11 B NMR spectroscopic analysis of the soluble decomposition products in CD3CN and D2O showed the formation of the protonated clusters [1-H-closo-1-CB11X11]− (X = H, F, Cl, Br), which is explained by the loss of elemental mercury. However, in an attempt to synthesize the so far unknown dianion [H11B11C−CB11H11]2−, in which both clusters are linked via a carbon−carbon bond, by heating of Cs21 to temperatures ranging from 250 to 400 °C in glass vessels under an Ar atmosphere or in a dynamic vacuum failed, and either unreacted starting material was recovered or insoluble polymeric material was obtained. The [Et4N]+ and Cs+ salts of [PhHg(closo-1-CB11X11)]− (X = H, Cl, Br, I (6−10)) are accessible from [closo-1-CB11X11]2− and PhHgCl (Scheme 1), similar to the synthesis of salts of the



RESULTS AND DISCUSSION Synthetic Aspects. The dianions [Hg(closo-1CB11X11)2]2− (X = H, F, Cl, Br (1−4)) are obtained by reactions of mercury dichloride and the dianions [closo-1CB11X11]2−, which were synthesized from the corresponding protonated cluster and n-butyllithium, as depicted in Scheme 1. The cesium and tetraethylammonium salts are precipitated from aqueous solutions, and in principle, the [Et4N]+ salts can be converted into the Cs+ salts via an extraction protocol as exemplified for the preparation of Cs21 in the Experimental Section. For the preparation of salts of 3 and 4, an alternative synthetic route was developed. Cs[1-H-closo-1-CB11X11] (X = Cl, Br) are reacted in situ with HgCl2 in methanol with KOH as a base (Scheme 2). This method is easier to perform and avoids inert conditions. However, it is not suitable for the synthesis of salts of 1 because [1-H-closo-1-CB11H11]− is not deprotonated in methanol/KOH. Because {closo-1-CB11F11} clusters undergo partial exchange of the fluorine substituents with hydroxy groups in basic aqueous solution,8,13,15 this method seems not to be suited for the preparation of salts of [Hg(closo-1CB11F11)2]2− (2), as well. 1567

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

Scheme 3. Attempted Synthesis of [C6F5Hg(closo-1-CB11X11)]− (X = H, F, Cl, Br, I)

Figure 1. (−)-ESI mass spectrum of [Hg(closo-1-CB11F11)2]2− (2) and the reaction products illustrating its fragmentation pathways. (a) CID of 2. (b) CID of the fragment [B10F10C]− (all isotopomers with m/z 310).

Scheme 4. Gas-Phase Reaction Pathways of Partially Fragmented Carborate Anions Generated by Reactions of [Hg(closo-1CB11X11)2]2− (X = F, C, Br, I (2−5)) in an Ion Trap

the fluorinated derivative 2 in Figure 1 (for the respective spectra of 3 and 4, see Figures S2 and S3 in the Supporting Information). In contrast, in the (−)-MALDI mass spectra, the corresponding radical anions [Hg(closo-1-CB11X11)2]−• were observed. Contrary to the formation of the dianions 1−4, the attempted syntheses of [Hg(closo-1-CB11I11)2]2− (5) failed by the reactions depicted in Schemes 1 and 2. The reaction of Cs[1-H-closo-1-CB11I11] with HgCl2 in basic methanol gave a mixture of singly and triply charged ions, but no evidence for the formation of 5 as a minor product was found (see Figure S4 in the Supporting Information). An unambiguous identification and structural assignment of the ions involved is complex and out of the scope of this study. However, the observation of

dianions 1−4. In contrast, the attempted preparation of salts of the anions [C6F5Hg(closo-1-CB11X11)]− (X = H, Cl, Br, I) starting from [closo-1-CB11X11]2− and C6F5HgCl failed, and the dianions 1−4 and Hg(C6F5)2 were obtained as main products (Scheme 3), as determined by NMR spectroscopy and (−)-ESI as well as (−)-MALDI mass spectrometry. In the reaction mixtures, trace amounts of [C6F5Hg(closo-1-CB11X11)]− (X = H, Cl, Br) were detected by mass spectrometry (see Figure S1 in the Supporting Information), whereas the corresponding anions with the perfluorinated and the periodinated cluster were not found, at all. Mass Spectrometry. In the (−)-ESI mass spectra, the complexes [Hg(closo-1-CB11X11)2]2− (X = H, F, Cl, Br (1−4)) were detected as dianions, as exemplified by the spectrum of 1568

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

I),18 and therefore, the presented mercury(II) complexes act in a certain way as a source of the dianions [closo-CB11X11]2− that are not accessible from [1-H-closo-1-CB11X11]− under these conditions.19 The fragmentation patterns of [PhHg(closo-1-CB11X11)]− (X = H, F, Cl, Br, I (6−10)) via CID followed by (−)-ESI mass spectrometric analysis are similar to the ones described for the dianions 1−5. The most important outcome of the study on 6− 10 is that, only for [PhHg(closo-1-CB11H11)]− (6), a further reaction was found that proceeds via elimination of Hg and formation of [1-Ph-closo-1-CB11H11]− under C−C bond formation (see Figure S8 in the Supporting Information). An analogous reaction was observed for the fragmentation of [C6F5Hg(closo-1-CB11H11)]−, which was formed as a very minor product and could be characterized by mass spectrometry, only. The formation of [1-C6F5-closo-1-CB11H11]− that also contains the nonhalogenated carba-closo-dodecaboranyl ligand is evident from the (−)-ESI mass spectrum in Figure S1 in the Supporting Information. A similar reaction was not found for the dianionic complex 1, which is composed of two [closo-1CB11H11]2− parts, probably because of the enhanced Coulomb repulsion of the two dianionic ligands. Crystal Structure Analysis. Cs2[Hg(closo-1-CB11X11)2] (X = H (Cs21), Cl (Cs23), Br (Cs24)) crystallize in the centrosymmetric space groups P21/c and P1̅ as the solvates Cs21·2Et2O, Cs23·MeCN, and Cs24·4Me2CO (Table 5). The donor atoms of the solvate molecules reveal weak interactions with the cesium cations (see Figures S9−S11 in the Supporting Information). No interaction between these donor atoms and the mercury atoms are present, at all. This observation is in strong contrast to the behavior of the homologous HgII complex with the fluorinated ligand [Hg(closo-1-CB11F11)2]2− (2), which crystallizes with either water or acetonitrile as a further ligand bonded to mercury.11 In all three structures, the mercury atom is located on a center of inversion, resulting in (i) linear C−Hg−C units and (ii) staggered orientations of the carba-closo-dodecaboranyl ligands (Figure 2). The Hg−C

trianions by (−)-ESI mass spectrometry extends the short list of experimentally observed comparatively small trianions that are stable in the gas phase.16 Collision-induced dissociation (CID) of these trianions in an ion trap yielded a broad variety of singly and doubly charged anions, including a species that, according to its mass, charge, and isotopic pattern, has the composition [C2B22HgI22]2−, as expected for the dianion 5 (see Figure S5 in the Supporting Information). The fragmentation reactions of the dianions 1−4 and [C2B22HgI22]2− were studied by CID in an ion trap, and the fragments were analyzed by (−)-ESI mass spectrometry. For the dianions with the halogen substituents Cl, Br, and I, analogous fragmentation pathways were observed, and the initial reaction is shown in eq 1. In the case of the reaction of 4, only the fragment [B11Br10C]− was present in the mass spectrum, whereas [BrHg(closo-1-CB11Br11)]− was not observed. The similar behavior of the dianions 3 and 4 as well as [C2B22HgI22]2− clearly shows that the latter dianion is [Hg(closo-1-CB11I11)2]2− (5).

[Hg(closo‐1‐CB11X11)2 ]2 − CID

⎯⎯⎯→ [XHg(closo‐1‐CB11X11)]− + [B11X10C]−

(1)



The ions [B11X10C] (X = Cl, Br, I) add water in the ion trap, followed by the loss of HX, to result in [B11X9OHC]−, as summarized in Scheme 4. Fragmentation of [XHg(closo-1CB11X11)]− (X = Cl, I) proceeds via loss of HgX2 to result in [B11X10C]− (X = Cl, I). For [ClHg(closo-1-CB11Cl11)]−, a competitive reaction via a homolytic cleavage of the Hg−C bond was observed that yields the radical anion [B11Cl11C]−•. [Hg(closo-1-CB11F11)2]2− (2) reveals a different behavior. Homolytic Hg−C bond cleavage occurs as the dominant fragmentation reaction (Figure 1). The formation of the radical fragment [B11F11C]−• may be explained by the strong B−F bond, which would have to be broken for the formation of [B11F10C]−. In summary, in the series of mercury(II) complexes with the halogenated ligands 2−5, formation of [B11X11C]−• is the major fragmentation reaction for X = F, formation of [B11X10C]− is the sole reaction found for X = Br and I, and both pathways were observed for X = Cl, although only for the fragmentation of [ClHg(closo-1-CB11Cl11)]−. Hence, this parallels the trend in B−Hal bond strengths, which decreases from fluorine to iodine (average bond dissociation energies of the boron trihalides: B−F 613, B−Cl 456, B−Br 357, B−I 260 kJ mol−1).17 The dianion [Hg(closo-1-CB11H11)2]2− (1) undergoes homolytic cleavage of the Hg−C bond similar to 2. For the perfluorinated cluster [B11F11C]−•, the degradation of the carborate “skeletal structure” occurs, whereas [B11Cl11C]−• does not show any degradation. Again, this might be the consequence of the strong B−F bond and the high stability of small neutral fragments, such as BF3. For [B11Cl10C]−, only a single loss of BCl3 was observed and higher energies were necessary and a much lower intensity of the fragmentation product was found. [B11X10C]− (X = Br, I) do not exhibit any degradation under B−B or B−C bond disruption under the reaction conditions used. The weakness of the B−I bond results in a fragmentation of [B11I10C]− that only proceeds via loss of iodine and finally yields the radical anion [B11IC]−• (see Figure S7 in the Supporting Information). These results show strong parallels to those derived from (−)-ESI mass spectrometric studies on the perhalogenated closo-dodecaborate dianions [closo-B12X12]2− (X = F, Cl, Br,

Figure 2. Dianions [Hg(closo-1-CB11X11)2]2− (X = H (left, 1), Cl (middle, 3), and Br (right, 4)) in the crystals of Cs21·2Et2O, Cs23·2MeCN, and Cs24·4Me2CO, respectively (displacement ellipsoids are at the 40% probability level). Selected bond lengths [Å]: Cs21·2Et2O, Hg1−C1 2.085(4); Cs23·2MeCN, Hg1−C1 2.080(6); Cs24·4Me2CO, Hg1−C1 2.085(10).

distances of 1, 3, and 4 are similar (2.080(6)−2.085(10) Å) and close to the values determined for [Et4N]2[Hg(closo-1CB11F11)2(NCMe)] ([Et4N]2[2·NCMe]) and [Et4N]4[{Hg1569

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

(closo-1-CB11F11)2}2(OH2)] ([Et4N]4[22·OH2]) (2.025(17)− 2.124(8) Å)11 and mercury(II) complexes with dicarba-closododecaboranyl ligands, for example, Hg(12-Ph-closo-1,12C2B10H10) (2.075 Å).4 [Et 4 N][PhHg(closo-1-CB 11 H 11 )]·0.5Me 2 CO ([Et 4 N] 6·0.5Me2CO) and Cs[PhHg(closo-1-CB11Br11)] (Cs9) crystallize in the monoclinic space groups P21 and P21/c, respectively. The acetone molecule in the structure of [Et4N]6 does not exhibit any interaction with the two independent anions 6, similar to the dianions 1, 3, and 4 in the structures of their Cs+ salts. The Hg−Ccluster distances given in Figure 3 are

and the halogen substituents bonded to the B5 belt that is connected to the cluster carbon atom cannot be fully excluded. The shortest intercluster Hal−Hal distances in 3 (4.244(4) Å) and 4 (4.198(3) Å) show that there is no significant repulsion between the two ligands coordinated to one mercury atom. The bonding properties of the mono- and dianionic HgII complexes calculated using different DFT methods and (RI)MP2 calculations are in good agreement with the experimental data (selected values are listed in Table S3 in the Supporting Information). At the slater-dirac-exchange and the S-VWN level of theory, the structure of [Hg(closo-1-CB11I11)2]2− (5) was calculated, as well. Slightly longer Hg−C bonds are predicted compared to the related mercury(II) complexes, which, in principle, is in agreement with the failure of its synthesis in the condensed phase. However, these differences are rather small, and hence, the failed preparation is more likely due to kinetic effects. This interpretation is supported by the successful generation of dianion 5 via CID, as described in the Mass Spectrometry section. Coordination of Acetonitrile and Water Solely to [Hg(closo-1-CB 11 F 11 ) 2 ] 2− and Not to [Hg(closo-1CB11X11)2]2− (X = H, Cl, Br)  A Theoretical Study. Only for the dianionic complex with the perfluorinated carbacloso-dodecaboranyl ligand was the coordination of a further ligand observed, so far.11 It is well documented in the literature that highly fluorinated {closo-1-CB11} clusters exhibit a distinct different chemistry compared with those with hydrogen or other halogen substituents.7,8,12−14,22 Hence, the unique coordination pattern observed for [Hg(closo-1-CB11F11)2]2− is a further example for such a different behavior. The coordination to mercury in organomercurials is mostly based on electrostatic interactions,5 and so the Lewis acidity of the mercury atom is one of the most important factors that determine the coordination chemistry of such organomercury compounds. A measure for the Lewis acidity of the mercury atoms in [Hg(closo-1-CB11X11)2]2− (X = H, F, Cl, Br, I (1−5)) is, to some degree, the proton affinity of [closo-1-CB11X11]2− and the reaction energies for the formations of [Hg(closo-1CB11X11)2]2− and [PhHg(closo-1-CB11X11)]−, which were calculated at different levels of theory (Table 1). For [closo-1CB11H11]2−, the highest proton affinity and the most exothermic reactions are predicted. Hence, this dianion is probably the best electron donor and the mercury atom in 1 should be the least Lewis acidic, which renders the coordination of a third ligand, such as CH3CN or H2O, unfavorable. The proton affinities and reaction energies within the series of [closo-1-CB11X11]2− dianions (X = halogen) are very similar, and for the fluorinated dianion, the majority of the calculations

Figure 3. Anions [PhHg(closo-1-CB11X11)]− (X = H (left, 6), Br (right, 9)) in the crystals of [Et4N]6·0.5Me2CO and Cs9, respectively (displacement ellipsoids are at the 40% probability level). Selected bond lengths [Å] and angles [deg]: [Et4N]6·0.5Me2CO, Hg1−C1 2.082(12), Hg1−C2 2.035(11), C1−Hg1−C2 174.7(5); Cs9, Hg1− C1 2.128(17), Hg1−C2 2.04(2), C1−Hg1−C2 175.9(7).

comparable to those found for 1−4 as well as for [Et4N][PhHg(closo-1-CB11F11)] ([Et4N]7) (2.111(12) Å).11 Slightly shorter d(Hg−C) are derived for the mercury−carbon bond with the phenyl ligand, in accordance with the findings reported for [Et4N]7 (2.062(11) Å). The small deviation from linearity of the C−Hg−C units of the complexes 6 and 9 as well as in 7 (174.2(4) °) is probably a result of packing effects. The shortest Hg−Br distances of 3.657(2) Å in the dianion 4 and of 3.606(2) Å in the single-negatively charged anion 9, as well as the shortest Hg−Cl distance of 3.514(4) Å in 3, are at the upper limit reported for the sum of the van der Waals radii of mercury (1.70−2.00 Å)20 and bromine (1.85 Å,21 in tert-alkyl groups) and, respectively, chlorine (1.77 Å,21 in tert-alkyl groups). Hence, weak interactions between the mercury atoms

Table 1. Calculated Energies of the Reactions of [closo-1-CB11X11]2− with H+, PhHgCl, and HgCl2a [closo-1-CB11X11]2− + H+ → [1-Hcloso-1-CB11X11]−

[closo-1-CB11X11]2− + PhHgCl → [PhHg(closo-1CB11X11)]− + Cl−

[closo-1-CB11X11]2− + 0.5HgCl2 → 0.5[Hg(closo-1-CB11X11)2]2− + Cl−

PAc

E

E

E

E

E

E

dianion

B3LYPb

MP2b

B3LYPb

MP2b

SDb

MP2b

SDb

[closo-1-CB11H11]2− [closo-1-CB11F11]2− [closo-1-CB11Cl11]2− [closo-1-CB11Br11]2− [closo-1-CB11I11]2−

1917.2 1682.4 1644.3 1622.6 1601.0

−1955.5 −1734.5 −1734.2 −1692.0 −1661.9

−426.8 −225.6 −183.2 −161.0 −140.4

−367.5 −185.2 −212.5 −178.3 −153.1

−361.1 −180.9 −156.1 −130.7 −108.7

−318.7 −138.3 −170.2 n.c.d n.c.d

−305.0 −129.3 −105.3 −79.1 −50.8

All energies in kJ mol−1. bMethods: B3LYP, B3LYP/6-311++G(d,p)/SDD; MP2, (RI)-MP2/def2-SVP/def2-ecp; SD, (RI)-slater-dirac/def2TZVPP/def2-ecp. cProton affinity (PA): PA = ΔH0g(base) + ΔH0g(H+) − ΔH0g(acid); ΔH0g(H+) = 2.5 RT. dn.c. = not calculated. a

1570

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

Table 2. Selected Experimental and Calculated Data of [Hg(closo-1-CB11X11)2(NCMe)]2− (X = H ({1·NCMe}), F ({2·NCMe}), Cl ({3·NCMe}))a,b species

methoda

{1·NCMe}c

MP2 S-VWN exptle

{2·NCMe}

{3·NCMe}

MeCN

MP2 SD S-VWN MP2 SD S-VWN exptlg MP2 SD S-VWN

d(Hg−C) [Å] 2.074/2.086 2.064/2.078 2.113(14)/ 2.025(17) 2.082/2.095 2.108/2.122 2.078/2.093 2.101/2.109 2.139/2.147 2.105/2.113

∠(C−Hg−C) [deg]

d(Hg···N) [Å]

d(CN) [Å]

∠(Hg···NC) [deg]

τb [deg]

ν(CN) [cm−1]

E [kJ mol−1]

175.5 175.4 167.9(6)

3.081 2.983 2.689(6)

1.177 1.157 1.123(10)

122.1 119.9 145.7(7)

2.8 0.4 0

n.c.d 2310 2271

−22.7 −19.9 n.d.f

172.1 171.4 171.3 169.7 168.9 168.6

2.770 2.699 2.583 2.742 2.670 2.568

1.175 1.167 1.155 1.174 1.166 1.154 1.157 1.175 1.168 1.156

133.5 132.0 132.3 139.7 142.9 141.6

1.9 5.1 5.0 25.8 25.9 26.3

n.c.d 2255 2328 n.c.d 2263 2333 2254 n.c.d 2251 2324

−27.2 −15.2 −25.6 5.2 10.2 4.5

a Methods: MP2, (RI)-MP2/def2-SVP/def2-ecp; SD, (RI)-slater-dirac/def2-TZVPP/def2-ecp; S-VWN, (RI)-S-VWN/def2-TZVPP/def2-ecp. bτ = deviation of the carboranyl ligands from an eclipsed conformation (0°). cNo minimum for the coordination of acetonitrile to the Hg atom of dianion 1 was found at the (RI)-slater-dirac/def2-TZVPP/def2-ecp level of theory. dn.c. = not calculated. e[Et4N]2[Hg(closo-1-CB11F11)2].11 fn.d. = not determined. gReferences 5 and 23.

Figure 4. Representations of the structures of the anionic complexes [Hg(closo-1-CB11X11)2(NCMe)]2− (X = H ({1·NCMe}), F ({2·NCMe}), Cl ({3·NCMe})) calculated at the (RI)-S-VWN/def2-TZVPP level of theory.

Table 3. Selected Experimental and Calculated Data of [Hg(closo-1-CB11F11)2(OH2)]2− ({1·OH2}) species {1·OH2}

H2O

methoda exptl

c

MP2 SD S-VWN exptlf SD S-VWN

d(Hg−C) [Å]

∠(C−Hg−C) [°]

d(Hg···O) [Å]

τb [°]

2.113(8)/2.087(9) 2.063(9)/2.124(8) 2.089/2.089 2.114/2.115 2.085/2.085

169.2(3) 171.3(3) 173.0 173.0 173.2

2.739(19) 2.85(2) 2.666 2.621 2.544

9.1(9) 2.8(9) 0.1 0.4 0.9

νas(OH) [cm−1]

νs(OH) [cm−1]

δ(HOH) [cm−1]

E [kJ mol−1]

3681

3583

1587

n.d.d

n.c.e 3506 3610 3756 3689 3835

n.c.e 3420 3527 3657 3580 3724

n.c.e 1500 1529 1595 1526 1554

−64.0 −53.3 −62.1

a Methods: MP2, (RI)-MP2/def2-SVP/def2-ecp; SD, (RI)-slater-dirac/def2-TZVPP/def2-ecp; S-VWN, (RI)-S-VWN/def2-TZVPP/def2-ecp. bτ = deviation of the carboranyl ligands from an eclipsed conformation (0°). c[Et4N]4[{Hg(closo-1-CB11F11)2}2(OH2)]: two different dianions 1 share one aqua. Top row: {1·OH2} with an occupancy of 70% for H2O. Bottom row: {1·OH2} with an occupancy of 30% for H2O.11 dn.d. = not determined. en.c. = not calculated. fGas phase.24

In Figure S12 in the Supporting Information, the space-filling models of the calculated structures of [Hg(closo-1-CB11X11)2]2− (X = H, F, Cl, Br, I (1−5)) at the (RI)-S-VWN level of theory are depicted. The steric shielding at mercury increases from X = H to I, and already for the chlorinated derivative, the central metal atom is effectively shielded, which prevents the interaction with a further ligand. The coordination of acetonitrile to [Hg(closo-1-CB11X11)2]2− (X = H, F, Cl) was modeled by DFT and (RI)-MP2 calculations. The results are

predict the most exothermic reactions and the highest proton affinity. This is, to some degree, surprising because the electronegativities suggest a different trend. However, these results indicate similar or even higher Lewis acidities for mercury in the higher homologues of [Hg(closo-1-CB11F11)2]2− (2) and do not explain the missing coordination of a third ligand to HgII in [Hg(closo-1-CB11X11)2]2− (X = Cl (3), Br (4)). 1571

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

NMR Spectroscopy. The 199Hg chemical shifts of the dianions [Hg(closo-1-CB11X11)2]2− (X = H, F, Cl, Br (1−4)) are found at lower resonance frequencies compared with δ(199Hg) of the respective anions [PhHg(closo-1-CB11X11)]2− (X = H, F, Cl (6−10)), as shown in Figure 6 and listed in Table

collected in Table 2, and models of the structures are plotted in Figure 4. The bond parameters calculated for [Hg(closo-1CB11F11)2(NCMe)]2− {2·NCMe} are in good agreement with experimental values; for example, the calculated structure is very close to an eclipsed conformation as present in the crystal structure. For {1·NCMe}, a very long d(Hg···N) was calculated, which suggests a weak interaction, only. However, for {2·NCMe} and {3·NCMe}, similar d(Hg···N) were predicted, indicating a similar strong coordination. However, for {3·NCMe}, the perchlorinated carba-closo-dodecaboranyl ligands are in between a staggered and an eclipsed conformation, showing the increased steric demand of these ligands. In agreement with this observation, coordination of acetonitrile to [Hg(closo-1-CB11Cl11)2]2− (3) is not favored, as shown by the calculated energy (Table 2). Hence, the most probable explanation for the nonappearance of the coordination of acetonitrile to 3 and also to 4 is the increased shielding of the central mercury atom rather than an electronic effect. The coordination of water to 2 was modeled at the different levels of theory used for the study on 2·NCMe, as well (Table 3 and Figure 5). The calculated structures of 2·OH2 are in close

Table 4. Selected NMR Spectroscopic Data of [Hg(closo-1CB11X11)2]2− (X = H, F, Cl, Br (1−4)) and [PhHg(closo-1CB11H11)]− (X = H, F, Cl, Br, I (6−10))a,b anion δ(199Hg) 1 2c 3 4 6 7c 8 9 10 a

81.9 43.3 80.1 85.7 87.3 47.7 85.2 93.0 103.0

δ(13C) Cphenyl

165.8 159.4 155.4 155.4 148.0

1

J(199Hg,13C) Ccluster

1

J(199Hg,13C) Cphenyl

1264 1702 1800 1930 1139 672 936 987 1092

1332 1994 1837 1844 1803

Solvent: CD3CN. bδ in ppm and J in Hz. cReference 11.

4. Within the two series, a decrease of δ(199Hg) is observed from the complex with the nonhalogenated cluster to the chlorinated derivative, followed by a small increase starting with the bromine-containing compounds 4 and 9, and even more for [PhHg(closo-1-CB11I11)]2− (10) in the series of phenyl complexes. In comparison to the parent carba-closo-dodecaborate anions [1-H-closo-1-CB11X11]−7 and the related anions [1-NC-closo-1CB11X11]− (X = H, F, Cl, Br, I),14 the δ(13C) of the cluster carbon atom is shifted to higher values for 1−4 and even more pronounced for 6−10. However, the trend in δ(13C) found within the four analogous series is similar (Table 4).14 The coupling constants between 199Hg and 13C of the cluster carbon atom are much larger for the homoleptic complexes than for the phenyl derivatives. In the row of the perhalogenated ligands, 1J(199Hg,13C) increases from fluorine to the heavier halogen atoms. For [PhHg(closo-1-CB11X11)]2− (X = H, F, Cl (6−10)), the increase in 1J(199Hg,13C) for the cluster carbon atom is paralleled by a decrease of 1J(199Hg,13C) of the ipsocarbon atom of the phenyl ligand.

Figure 5. Two representations of the structure of the aqua complex [Hg(closo-1-CB11F11)2(OH2)]2− ({2·OH2}) calculated at the (RI)-SVWN/def2-TZVPP level of theory.

agreement with those of [Et4N]4[{Hg(closo-1CB11F11)2}2(OH2)] ([Et4N]4[22·OH2]),11 where two dianions 2 share a single water molecule. The agreement is especially good with the dianion that is bonded to the water molecule with an occupancy of 70%, which has, therefore, a higher percentage of 2·OH2 than of 2.

Figure 6.

−1018 −1064 −1340 −1311 −917 −975 −1057 −1002 −865

δ(13C) Ccluster

199

Hg{1H} NMR spectra of [Hg(closo-1-CB11X11)2]2− (X = H, F, Cl (1−3)) and [PhHg(closo-1-CB11X11)]2− (X = H, F, Cl (6−8)). 1572

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

Table 5. Selected Crystal Data and Details of the Refinement of the Crystal Structures of Cs2[Hg(closo-1-CB11H11)2]·2Et2O (Cs21·2Et2O), Cs2[Hg(closo-1-CB11Cl11)2]·MeCN (Cs23·MeCN), Cs2[Hg(closo-1-CB11Br11)2]·4Me2CO (Cs24·4Me2CO), [Et4N][PhHg(closo-1-CB11H11)]·0.5Me2CO ([Et4N]6·0.5Me2CO), and Cs[PhHg(closo-1-CB11Br11)] (Cs9) empirical formula formula wt T (K) cryst syst space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) volume (Å3) Z Dcalcd (mg m−3) μ (mm−1) F(000) no. of collected reflns no. of unique reflns, R(int) no. of parameters/restraints R1 (I > 2σ(I)) wR2 (all) GOF on F2 largest diff. peak/hole (e Å−3) CCDC no.



Cs21·2Et2O

Cs23·MeCN

Cs24·4Me2CO

[Et4N]6·0.5Me2CO

Cs9

C10H42B22Cs2HgO2 898.67 173 monoclinic P21/c 13.1322(8) 11.4190(5) 11.7541(6)

C4H3B22Cl22Cs2HgN 1549.20 173 triclinic P1̅ 9.145(2) 9.439(2) 14.881(3) 79.62(3) 73.49(3) 66.73(3) 1128.0(4) 1 2.281 6.320 708 9681 3930, 0.050 244/3 0.042 0.128 1.127 1.420/−2.041 847882

C14H24B22Br22Cs2HgO4 2718.58 290 triclinic P1̅ 11.853(2) 12.584(3) 12.958(3) 70.20(4) 66.00(3) 65.59(3) 1573.7(8) 1 2.869 17.577 1210 13 939 5478, 0.048 299/0 0.044 0.119 1.045 1.753/−1.472 847881

C33H78B22Hg2N2O 1157.97 295 monoclinic P21 10.908(2) 17.524(4) 13.362(3)

C7H5B11Br11CsHg 1420.53 290 monoclinic P21/c 15.261(3) 9.556(2) 20.195(4)

101.36(3)

108.83(3)

2504.2(9) 2 1.536 6.153 1136 25 896 7392, 0.050 551/145 0.037 0.118 1.082 1.888/−1.794 847885

2787.5(10) 4 3.385 22.577 2488 22 790 5998, 0.128 280/30 0.082 0.198 1.190 1.534/−2.771 847884

115.264(7) 1594.01(17) 2 1.872 7.090 836 9157 2798, 0.029 226/68 0.019 0.040 1.044 0.796/−0.841 847883

closo-dodecaborate dianions.25 Furthermore, mercury complexes of this type with carba-closo-dodecaboranyl ligands that have functional groups bonded to the cluster boron atoms, for example, alkynyl substituents, are potential building blocks for coordination polymers and supramolecular chemistry.

SUMMARY AND CONCLUSION The homoleptic mercury(II) complexes of carba-closo-dodecaboranyl ligands [Hg(closo-1-CB11X11)2]2− (X = H, F, Cl, Br (1− 5)) are easily accessible. As a result of the high steric demand of the [closo-1-CB11I11]2− ligand, the coordination of two of these ligands to mercury(II) failed in the condensed phase. However, in principle, two of these bulky ligands can bind to one mercury center, as proven by collision-induced fragmentation reactions starting from trianions containing three {Hg(closo-1-CB11I11)} segments, which resulted in the formation of [Hg(closo-1CB11I11)2]2− (5) in the gas phase. The synthesis of phenylmercury complexes with one [closo-1-CB11X11]2− ligand [PhHg(closo-1-CB11X11)]− (X = H, F, Cl, Br, I (6−10)) is straightforward, whereas for the attempted synthesis of [C6F5Hg(closo-1-CB11X11)]− (X = H, F, Cl, Br, I), dismutation was observed to result in [Hg(closo-1-CB11X11)2]2− and Hg(C6F5)2. Only for [Hg(closo-1-CB11F11)2]2− (2) was the coordination of a third, neutral ligand to mercury observed,11 whereas for its higher homologues 3 and 4 and the nonhalogenated anion 1, no additional coordination was found. In the case of dianion 1, the reduced Lewis acidity of the mercury atom is most likely the reason for the different behavior. In contrast, for the chlorinated and brominated derivatives 3 and 4, the Lewis acidity of mercury is probably similar or slightly higher compared to that of the mercury center in 2. The increased steric demand of the ligands as a result of the larger halogen substituents prevents in the case of 3 and 4 the interaction with a further ligand. Because of their high thermal and chemical stability, the dianionic homoleptic complexes with the perhalogenated ligands [Hg(closo-1-CB11X11)2]2− (X = F, Cl, Br (2−4)) may serve as weakly coordinating dianions similar to halogenated



EXPERIMENTAL SECTION

General Methods. 1H, 11B, 13C, 19F, and 199Hg NMR spectra were recorded at 25 °C in CD3CN on a Bruker Avance III 400 spectrometer operating at 400.17 (1H), 128.39 (11B), 100.62 (13C), 376.45 (19F), and 71.67 MHz (199Hg) or on a Bruker Avance DRX500 spectrometer operating at 500.13 (1H), 125.76 (13C), 160.46 (11B), 470.59 (19F), and 89.58 MHz (199Hg). The NMR signals were referenced against TMS (1H, 13C), BF3·OEt2 in CD3CN (11B), CFCl3 (19F), and Ph2Hg in DMSO-d6 (1 mol L−1) (199Hg, δ(199Hg) = −1187 ppm;26 Me2Hg, δ(199Hg) = 0 ppm) as external standards. Elemental analyses (C, H, N) were performed with a Euro EA3000 instrument (HEKA-Tech, Germany). Thermoanalytical measurements were made with a Mettler Toledo DSC 30 instrument. Temperature and sensitivity calibrations in the temperature range of 25−600 °C were carried out on samples of Ga, In, Pb, Sn, and Zn. About 5−10 mg of the solid samples was weighed and contained in sealed aluminum crucibles, and the studies were performed with a heating rate of 5 K min−1. Throughout this process, the furnace was flushed with dry nitrogen. Chemicals. All standard chemicals were obtained from commercial sources. Cesium carba-closo-dodecaborate was synthesized from [Me3NH][nido-B11H14]27 following an improved protocol28 of a literature procedure.29 The salts Cs[1-H-closo-1-CB11Cl11],30 Cs[1-Hcloso-1-CB11Br11],30 and Cs[1-H-closo-1-CB11I11]30 were prepared similar to published procedures starting from Cs[closo-1-CB11H12]. [Et4N]2[Hg(closo-1-CB11F11)2] and [Et4N][PhHg(closo-1-CB11F11)] were synthesized as described recently.11 The synthesis of C6F5HgCl is described in the Supporting Information. Mass Spectrometry. ESI-MS measurements were performed on a Bruker Esquire-LC ion trap mass spectrometer (Bruker Daltonik, 1573

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

small-core potential was used (def2-ecp47). All structures of the DFT calculations represent true minima with no imaginary frequency on the respective hypersurface. For the (RI)-MP2 calculations, no frequency analyses were performed. Synthesis of [Et4N]2[Hg(closo-1-CB11H11)2] ([Et4N]21). A 70 mL glass finger equipped with a valve with a PTFE stem (Young, London), fitted with a PTFE-coated magnetic stirring bar, was charged with Cs[1-H-closo-1-CB11H11] (300 mg, 1.10 mmol) and tetrahydrofuran (15 mL). After addition of a solution of nBuLi in hexane (0.68 mL, 1.6 mol L−1, 1.10 mmol) at 0 °C, the reaction mixture was warmed to room temperature and stirred for 1 h. A solution of HgCl2 (152 mg, 0.55 mmol) in THF (15 mL) was added, and the resulting mixture was stirred for further 2 h. The reaction mixture was poured into water (50 mL), and the THF was removed using a rotary evaporator. An aqueous solution of [Et4N]Br (1.50 g, 7.14 mmol, 10 mL) was added slowly, resulting in the immediate formation of a white precipitate. This solid was filtered off and dried in a vacuum. The crude product, which contained a small amount of [Et4N][1-H-closo-1CB11H11] as an impurity, was dissolved in acetone (10 mL), filtered, precipitated by the addition of diethyl ether (50 mL), and dried in a vacuum to yield pure [Et4N]21. Yield: 286 mg (0.38 mmol, 70%). 1 H{11B} NMR (CD3CN, δ ppm): 1.51 (s, 10H, BH7−11), 1.50 (s, 10H, 3J(199Hg,1H) = 61.6 Hz, BH2−6), 1.17 (s, 2H, BH12). 13C{1H} NMR (CD3CN, δ ppm): 81.9 (s, 1J(199Hg,13C) = 1250 Hz, Ccluster). 11 B NMR (CD3CN, δ ppm): −4.9 (d, 2B, 1J(11B,1H) = 132 Hz, B12), −11.8 (d, 10B, 1J(11B,1H) = 136 Hz, B7−11), −14.2 (d, 10B, 1 11 1 J( B, H) = 148 Hz, B2−6). (−)-MALDI-MS m/z (isotopic abundance > 60) calcd for 1 ([C2H22B22Hg]−): 483(69), 484(77), 485(89), 486(100), 487(85). Found: 483(71), 484(79), 485(90), 486(100), 487(84). Anal. Calcd for C18H62B22HgN2: C, 29.01; H, 8.39; N, 3.76. Found: C, 29.24; H, 8.42; N, 3.76. [Et4N]2[Hg(closo-1-13CB11H11)2]. The 13C-labeled compound was synthesized as described for the nonisotopically enriched compounds. 199 Hg{1H} NMR (CD3CN, δ ppm): −1018 (s, 1J(199Hg,13C) = 1264 Hz). Preparation of Cs2[Hg(closo-1-CB11H11)2] (Cs21). In a 250 mL Erlenmeyer flask, [Et4N]21 (160 mg, 0.21 mmol) was suspended in 10 mL of hydrochloric acid (10% v/v), and 100 mL of diethyl ether was added. The mixture was stirred until all solid material dissolved. The ether layer was separated, and the aqueous phase was extracted two more times with Et2O (2 × 50 mL). The combined ether solutions that contain dianion 1 and H+(solv) as the countercation were dried with MgSO4 and filtered, and the volume of the solution was reduced to approximately 20 mL. The resulting mixture was treated with a solution of CsCl (200 mg, 1.19 mmol) in water (5 mL). Ether and water were removed under reduced pressure, and the colorless solid residue was extracted with a total of 60 mL of acetone. The acetone was evaporated, and the semisolid residue was treated with 100 mL of CHCl3. The mixture was stored in a refrigerator for 2 h. The white solid was isolated by filtration. Yield: 134 mg (0.18 mmol, 85%). Anal. Calcd for C2H22B22Cs2Hg: C, 3.20; H, 2.95. Found: C, 3.35; H, 2.88. Synthesis of Cs 2 [Hg(closo-1-CB 11 Cl 11 ) 2 ] (Cs 2 3) and [Et4N]2[Hg(closo-1-CB11Cl11)2] ([Et4N]23) Method A. The preparation was performed as described for [Et4N]21 starting from Cs[1H-closo-1-CB11Cl11] (500 mg, 0.76 mmol) and HgCl2 (104 mg, 0.38 mmol) except for the isolation of the salt(s). The crude product was dissolved in water (25 mL), and the addition of an aqueous solution of CsCl (2 g, 11.9 mmol, 10 mL) resulted in the formation of Cs23. After filtration, the crude Cs23 was dissolved in acetone (10 mL) and filtered, and after removal of most of the acetone, pure Cs23 was obtained as a white solid by the addition of chloroform. Yield: 316 mg (0.21 mmol, 55%). The remaining aqueous solution after the precipitation of the crude cesium salt was diluted with water (25 mL), and slow addition of a solution of [Et4N]Br (1.50 g, 7.14 mmol) in water (30 mL) resulted in the formation of a white solid that was contaminated with a small amount of [Et4N][1-H-closo-1-CB11Cl11]. Pure [Et4N]23 was obtained via recrystallization from acetone by slow uptake of Et2O vapor. Yield: 206 mg (0.14 mmol, 16%). NMR data for anion 3: 13C{1H} NMR (CD3CN, δ ppm): 80.1 (s, 1J(199Hg,13C) = 1800 Hz, Ccluster). 11B NMR (CD3CN, δ ppm): −3.0 (s, 2B, B12),

Bremen, Germany,). Samples were dissolved in acetonitrile (LCMS grade; VWR, Darmstadt, Germany) at concentrations of approximately 10−5−10−6 mol L−1 and injected into the mass spectrometer via a syringe pump at a flow rate of 3 μL min−1. Spectra were recorded in the negative ion mode for 3−5 min and averaged. Collision-induced dissociation (CID) was induced in an ion trap. Ions with a defined m/z value are isolated and specifically activated by collisions with the background gas (mainly He, and some N2 and H2O) in the ion trap. The collision is induced by applying an adjustable ac voltage to the ion trap end caps. The resulting mass spectrum shows only the reaction products of the isolated and activated ions. All fragmentation steps were verified by tandem-MS experiments. Matrix-assisted laser desorption/ionization (MALDI) mass spectra in the negative-ion mode were recorded on a Bruker Ultraflex TOF spectrometer. Single-Crystal X-ray Diffraction. Colorless crystals of Cs21·2Et2O suitable for X-ray diffraction studies were grown from a mixture of diethyl ether and acetone (1:1 v/v) by slow evaporation of the solvents. Similarly, crystals of Cs23·MeCN and Cs24·4Me2CO were obtained from acetonitrile and acetone, respectively. Slow uptake of diethyl ether into a solution of [Et4N]6·0.5Me2CO in acetone resulted in colorless crystals, and crystals of Cs9 were grown from acetone by slow evaporation of the solvent. A crystal of Cs21·2Et2O, Cs23·MeCN, and [Et4N]6·0.5Me2CO was investigated with an imaging plate diffraction system (IPDS, Stoe & Cie), and a crystal of Cs24·4Me2CO and Cs9 was studied using a Stoe STADI CCD diffractometer using Mo Kα radiation (λ = 0.71073 Å). Numerical absorption corrections31 based on indexed crystal faces were applied to the data of all compounds after optimization of the crystal shape.32 All structures were solved by direct methods,33,34 and refinement is based on full-matrix least-squares calculations on F2.34,35 The positions of most of the hydrogen atoms in the crystal structures were located via ΔF syntheses. The only exceptions are those of the hydrogen atoms of the acetone molecules in [Et4N] 6·0.5Me2CO and Cs24·4Me2CO as well as of the disordered solvent molecules Et2O and acetonitrile in Cs21·2Et2O and Cs23·MeCN, respectively. All non-hydrogen atoms were refined anisotropically. Because of the disorder of Et2O and MeCN in the structures of the cesium salts of 1 and 3, distance and similarity restraints for the anisotropic displacement parameters of these solvent molecules were necessary. In the structures of [Et4N]6·0.5Me2CO and Cs9, similarity restraints for the anisotropic displacement parameters of some of the cluster carbon and boron atoms were introduced to achieve a stable refinement. All hydrogen atoms were refined using idealized bond lengths as well as angles, and their isotropic displacement parameters were kept equal to 120% of the respective parent carbon or boron atom except for the hydrogen atoms of the CH3 groups, whose isotropic displacement parameters were set to 150% of the respective parent carbon atom. Molecular structure diagrams were drawn with the program Diamond 3.2g.36 Experimental details, crystal data, and CCDC numbers are collected in Table 5. The supplementary crystallographic data for this publication are found in the Supporting Information or can be obtained free of charge from the Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif. Quantum Chemical Calculations. Density functional calculations (DFT)37 using the hybrid functional B3LYP38 and Pople-type basis sets 6-311++G(d,p)39 for all atoms except for iodine were performed with the Gaussian 03 program suite.40 For iodine and mercury, pseudopotentials were used as implemented in the SDD basis sets in the Gaussian 03 program suite.40 DFT calculations with the LDA functionals slater-dirac-exchange,41 and S-VWN,41,42 as well as ab initio calculations employing second-order Møller−Plesset (MP2) perturbation theory43 were done with the Turbomole V6.0 program package.44 The resolution-of-the-identity approximation was used for S-VWN and slater-dirac-exchange ((RI)-DFT45) as well as MP2 calculations ((RI)-CC246). def2-TZVPP basis sets were employed for (RI)-S-VWN and (RI)-slater-dirac-exchange, and def2-SVP basis sets for (RI)-MP2 calculations.47 For iodine and mercury, a relativistic 1574

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

−8.9 (s, 10B, B7−11), −11.4 (s, 10B, B2−6). 199Hg{1H} NMR (CD3CN, δ ppm): −1340 (s). (−)-MALDI-MS m/z (isotopic abundance > 60) calcd for 3 ([C 2 B 22 Cl22Hg] − ): 1497(61), 1499(81), 1501(85), 1502(81), 1503(100), 1504(71), 1505(84). Found: 1497(65), 1499(80), 1501(80), 1502(78), 1503(100), 1504(75), 1505(86). Cs23: Anal. Calcd for C2B22Cl22Cs2Hg: C, 1.59. Found: C, 1.51. [Et4N]23: Anal. Calcd for C18H40B22Cl22HgN2: C, 14.39; H, 2.68; N, 1.86. Found: C, 15.06; H, 3.12; N, 1.51. Synthesis of Cs2[Hg(closo-1-CB11Cl11)2] (Cs23). Method B. Cs[1-H-closo-1-CB11Cl11] (260 mg, 0.4 mmol) and 5 equiv of KOH were dissolved in methanol (10 mL), and subsequently, HgCl2 (54 mg, 0.2 mmol) was added to the clear solution. The reaction mixture was refluxed for 12 h. After cooling to room temperature, deionized water (50 mL) was added and most of the methanol was removed at a rotary evaporator. Dropwise addition of a solution of CsCl (2.00 g, 11.88 mmol) in water (10 mL) resulted in the precipitation of pure Cs23 as a colorless solid. Yield: 226 mg (0.15 mmol, 76%). Preparation of Cs2[Hg(closo-1-CB11Br11)2] (Cs24). The synthesis of Cs24 was accomplished by both methods described for the preparation of Cs23. Method A. Cs[1-H-closo-1-CB11Br11] (300 mg, 0.26 mmol) was treated with nBuLi in hexane (0.18 mL, 1.6 mol L−1, 0.28 mmol) at 0 °C, followed by the addition of HgCl2 (35 mg, 0.13 mmol). Yield: 270 mg (0.11 mmol, 83%). Method B. Cs[1-H-closo-1-CB11Br11] (550 mg, 0.48 mmol) was reacted with HgCl2 (65 mg, 0.24 mmol) in basic methanol. Yield: 507 mg (0.20 mmol, 85%). 13C{1H} NMR (CD3CN, δ ppm): 85.7 (s, 1 199 J( Hg,13C) = 1930 Hz, Ccluster). 11B NMR (CD3CN, δ ppm): −3.8 (s, 2B, B12), −9.2 (s, 10B, B7−11), −11.6 (s, 10B, B2−6). 199Hg NMR (CD3CN, δ ppm): −1311 (s). (−)-MALDI-MS m/z (isotopic abundance > 60) calcd for 4 ([C2B22Br22Hg]−): 2216(65), 2217(84), 2218(84), 2219(95), 2220(98), 2221(90), 2222(100), 2223(84), 2224(76), 2225(68), 2226(66). Found: 2216(60), 2217(82), 2218(80), 2219(95), 2220(95), 2221(86), 2222(100), 2223(89), 2224(80), 2225(75), 2226(69). Anal. Calcd for C2B22Br22Cs2Hg: C, 0.97. Found: C, 1.35. Preparation of [Et4N][PhHg(closo-1-13CB11H11)] ([Et4N]6). A 70 mL glass finger equipped with a valve with a PTFE stem (Young, London), fitted with a PTFE-coated magnetic stirring bar, was charged with Cs[1-H-closo-1-13CB11H11] (100 mg, 0.36 mmol) and tetrahydrofuran (20 mL). After addition of a solution of nBuLi in hexane (0.25 mL, 1.6 mol L−1, 0.40 mmol) at 0 °C, the reaction mixture was warmed to room temperature and stirred for 1 h. A solution of PhHgCl (126 mg, 0.40 mmol) in THF (15 mL) was added, and the resulting mixture was stirred for a further 2 h. The reaction mixture was poured into water (50 mL), the THF was removed using a rotary evaporator, and the aqueous phase was filtered. A solution of [Et4N]Br (1.50 g, 7.14 mmol) in H2O (30 mL) was added slowly under stirring. The white precipitate that had formed was isolated by filtration. The crude product that contained a small amount of Cs[1-H-closo1-13CB11H11] was dried in a vacuum and dissolved in acetonitrile, and the solution was filtered. The volume of the clear acetonitrile solution was reduced to approximately 5 mL, and addition of diethyl ether (100 mL) resulted in the precipitation of [Et4N]6 as a white solid, which was filtered off, and dried in a vacuum. Yield: 169 mg (0.31 mmol, 85%). NMR data for anion 6: 1H{11B} NMR (CD3CN, δ ppm): 7.5− 7.1 (m, 5H, phenyl), 1.69 (s, 5H, 3J(199Hg,1H) = 44 Hz, BH2−6), 1.59 (s, 5H, BH7−11), 1.21 (s, 1H, BH12). 13C{1H} NMR (CD3CN, δ ppm): 165.8 (d, 1C, 2J(13C,13C) = 37.4 Hz, 1J(199Hg,13C) = 1332 Hz, Cipso), 138.8 (s, 2C, 2J(199Hg,13C) = 92 Hz, Cortho), 129.1 (d, 2C, 4 13 13 J( C, C) = 2.0 Hz, 3J(199Hg,13C) = 112 Hz, Cmeta), 128.9 (s, 1C, 4 199 J( Hg,13C) = 19 Hz, Cpara), 87.3 (s, 1C, 1J(199Hg,13C) = 1150 Hz, 2 13 13 J( C, C) = 37 Hz, Ccluster). 11B NMR (CD3CN, δ ppm): −5.0 (d, 1B, 1J(11B,1H) = 133 Hz, B12), −11.5 (d, 5B, 1J(11B,1H) = 135 Hz, B7−11), −13.8 (d, 5B, 1J(11B,1H) = 149 Hz, B2−6). 199Hg{1H} NMR (CD3CN, δ ppm): −917 (s, 1J(199Hg,13C) = 1139 Hz). (−)-MALDIMS m/z (isotopic abundance > 60) calcd for 6 ([C7H16B11Hg]−): 419(83), 420(100), 421(100), 422(84). Found: 419(85), 420(97),

421(100), 422(81). Anal. Calcd for C15H36B11HgN: C, 32.76; H, 6.60; N, 2.55. Found: C, 32.91; H, 6.72; N, 2.35. Synthesis of Cs[PhHg(closo-1-CB11Cl11)] (Cs8). The synthesis of Cs8 was performed similarly to the preparation of [Et4N]6 starting from Cs[1-H-closo-1-CB11Cl11] (200 mg, 0.38 mmol) and PhHgCl (131 mg, 0.42 mmol). However, a solution of CsCl (2 g, 11.88 mmol) in water (5 mL) was slowly added to an aqueous solution of the crude product, resulting in the precipitation of an off-white solid. Crude Cs8 was dissolved in acetone and filtered. Chloroform (50 mL) was added, and the acetone was removed at a rotary evaporator. Colorless, pure Cs8 precipitated from the solution at 5 °C overnight. Yield: 196 mg (0.21 mmol, 55%). 1H{11B} NMR (CD3CN, δ ppm): 7.5−7.0 (m, phenyl). 13C{1H} NMR (CD3CN, δ ppm): 155.4 (s, 1C, 1J(199Hg,13C) = 1837 Hz, Cipso), 137.3 (s, 2C, 2J(199Hg,13C) = 108 Hz, Cortho), 130.4 (s, 1C, 4J(199Hg,13C) = 28 Hz, Cpara), 129.9 (s, 2C, 3J(199Hg,13C) = 156 Hz, Cmeta), 85.2 (s, 1C, 1J(199Hg,13C) = 936 Hz, Ccluster). 11B NMR (CD3CN, δ ppm): −3.8 (s, 1B, B12), −9.2 (s, 5B, B7−11), −11.6 (s, 5B, B2−6). 199Hg{1H} NMR (CD3CN, δ ppm): −1057 (s). (−)-MALDI-MS m/z (isotopic abundance > 60) calcd for 8 ([C 7 H 5 B 11 Cl 11 Hg] − ): 795(89), 796(92), 797(80), 798(100), 799(65), 800(87). Found: 795(86), 796(95), 797(75), 798(100), 799(69), 800(91). Anal. Calcd for C7H5B11Cl11CsHg: C, 9.03; H, 0.54. Found: C, 9.95; H, 0.57. Synthesis of Cs[PhHg(closo-1-CB11Br11)] (Cs9). Cs9 was prepared as described for Cs8 using Cs[1-H-closo-1-CB11Br11] (250 mg, 0.22 mmol) and PhHgCl (75 mg, 0.24 mmol) as starting materials. Yield: 240 mg (0.17 mmol, 76%). 1H{11B} NMR (CD3CN, δ ppm): 7.4−7.0 (m, phenyl). 13C{1H} NMR (CD3CN, δ ppm): 155.4 (s, 1C, 1J(199Hg,13C) = 1844 Hz, Cipso), 137.2 (s, 2C, 2J(199Hg,13C) = 109 Hz, Cortho), 130.5 (s, 1C, 4J(199Hg,13C) = 28 Hz, Cpara), 129.9 (s, 2C, 3J(199Hg,13C) = 155 Hz, Cmeta), 93.0 (s, 1C, 1J(199Hg,13C) = 987 Hz, Ccluster). 11B NMR (CD3CN, δ ppm): −4.3 (s, 1B, B12), −8.6 (s, 5B, B7−11), −11.8 (s, 5B, B2−6). 199Hg{1H} NMR (CD3CN, δ ppm): −1002 (s). (−)-MALDI-MS m/z (isotopic abundance > 60) calcd for 9 ([C7H5B11Br11Hg]−): 1283(64), 1284(68), 1285(62), 1286(100), 1287(97), 1288(99), 1289(83), 1290(94), 1291(76), 1292(65). Found: 1283(65), 1284(71), 1285(66), 1286(100), 1287(99), 1288(100), 1289(85), 1290(97), 1291(77), 1292(64). Anal. Calcd for C7H5B11Br11CsHg: C, 5.92; H, 0.35. Found: C, 5.85; H, 0.36. Synthesis of Cs[PhHg(closo-1-CB11I11)] (Cs10). In analogy to the synthesis described for Cs8, the cesium salt of anion 10 was prepared from Cs[1-H-closo-1-CB11I11] (300 mg, 0.18 mmol) and PhHgCl (63 mg, 0.20 mmol). Yield: 290 mg (0.15 mmol, 83%). 1 H{11B} NMR (CD3CN, δ ppm): 7.4−7.0 (m, phenyl). 13C{1H} NMR (CD3CN, δ ppm): 148.0 (s, 1C, 1J(199Hg,13C) = 1803 Hz, Cipso), 137.3 (s, 2C, 2J(199Hg,13C) = 105 Hz, Cortho), 130.2 (s, 1C, 4 199 J( Hg,13C) ∼ 22 Hz, Cpara), 129.5 (s, 2C, 3J(199Hg,13C) = 153 Hz, Cmeta), 103.0 (s, 1C, 1J(199Hg,13C) = 1092 Hz, Ccluster). 11B NMR (CD3CN, δ ppm): −9.9 (s, 1B, B12), −11.4 (s, 5B, B7−11), −16.5 (s, 5B, B2−6). 199Hg{1H} NMR (CD3CN, δ ppm): −865 (s). (−)-MALDI-MS m/z (isotopic abundance > 60) calcd for 10 ([C7H5B11HgI11]−): 1804(77), 1805(74), 1806(100), 1807(80). Found: 1804(80), 1805(77), 1806(100), 1807(82). Anal. Calcd for C7H5B11CsHgI11: C, 4.34; H, 0.26. Found: C, 4.68; H, 0.38. Attempted Synthesis of Cs[C6F5Hg(closo-1-CB11X11)] (X = H, F, Cl, Br, I). The reactions were performed similar to the syntheses of the Cs+ and [Et4N]+ salts of the anions 6 and 8−10. The solution of C6F5HgCl in THF was added either at 0 °C or at −78 °C to a solution of the corresponding Li+/Cs+ salt of the carboranyl ligand in THF. The NMR spectroscopic and mass spectrometric ((−)-ESI and (−)-MALDI) analyses showed no differences for the reactions performed at the two different temperatures. According to the 11B and 19F NMR spectroscopic data of the reaction mixtures, the main products (>90%) of the reactions are Hg(C6F5)2 and the dianions 1−4 and a complex mixture of {closo-1-CB11I11} derivatives, which probably contain some carba-closo-dodecaboranyl mercury(II) complexes, in the case of the iodinated carba-closo-dodecaborate cluster, respectively. 1575

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics



Article

(13) Finze, M.; Reiss, G. J.; Zähres, M. Inorg. Chem. 2007, 46, 9873− 9883. (14) Sprenger, J. A. P.; Finze, M.; Schaack, B. B. Dalton Trans. 2010, 39, 2708−2716. (15) Strauss, S. H.; Ivanov, S. V.; Lupinetti, A. J. Patent Number WO 98/43983; Colorado State University Research Foundation, 1998. (16) Sommerfeld, T.; Bhattarai, B. Phys. Chem. Chem. Phys. 2011, 13, 18393−18397. (17) (a) Koski, W. S.; Kaufman, J. J.; Pachucki, C. F. J. Am. Chem. Soc. 1959, 81, 1326−1331. (b) Huheey, J. E.; Keiter, E.; Keiter, R. L. Anorganische Chemie. Prinzipien von Struktur und Reaktivität., 3rd ed.; de Gruyter: Berlin, 2003. (18) Warneke, J.; Dülcks, T.; Knapp, C.; Gabel, D. Phys. Chem. Chem. Phys. 2011, 13, 5712−5721. (19) Warneke, J.; Himmelspach, A.; Finze, M. Unpublished results. (20) Canty, A. J.; Deacon, G. B. Inorg. Chim. Acta 1980, 45, L225− L227. (21) Bondi, A. J. Phys. Chem. 1964, 68, 441−451. (22) Kobayashi, Y.; Popov, A. A.; Miller, S. M.; Anderson, O. P.; Strauss, S. H. Inorg. Chem. 2007, 46, 8505−8507. (23) Karakida, K.; Fukuyama, T.; Kuchitsu, K. Bull. Chem. Soc. Jpn. 1974, 47, 299. (24) Benedict, W. S.; Gailar, N.; Plyler, E. K. J. Chem. Phys. 1956, 24, 1139−1165. (25) (a) Ivanov, S. V.; Miller, S. M.; Anderson, O. P.; Solntsev, K. A.; Strauss, S. H. J. Am. Chem. Soc. 2003, 125, 4694−4695. (b) Avelar, A.; Tham, F. S.; Reed, C. A. Angew. Chem. 2009, 121, 3543−3545;(c) Angew. Chem., Int. Ed. 2009, 48, 3491−3493. (d) Kessler, M.; Knapp, C.; Sagawe, V.; Scherer, H.; Uzun, R. Inorg. Chem. 2010, 122, 5223− 5230. (26) Sen, M. A.; Wilson, N. K.; Ellis, P. D.; Odom, J. D. J. Magn. Reson. 1975, 19, 323−336. (27) Dunks, G. B.; Barker, K.; Hedaya, E.; Hefner, C.; Palmerordonez, K.; Remec, P. Inorg. Chem. 1981, 20, 1692−1697. (28) (a) Küppers, T.; Bernhardt, E.; Willner, H. Angew. Chem. 2007, 119, 6462−6465;(b) Angew. Chem., Int. Ed. 2007, 46, 6346−6349. (29) Franken, A.; King, B. T.; Rudolph, J.; Rao, P.; Noll, B. C.; Michl, J. Collect. Czech. Chem. Commun. 2001, 66, 1238−1249. (30) Xie, Z.; Tsang, C.-W.; Sze, E. T.-P.; Yang, Q.; Chan, D. T. W.; Mak, T. C. W. Inorg. Chem. 1998, 37, 6444−6451. (31) X-RED 1.20: Stoe Data Reduction Program; STOE & Cie GmbH: Darmstadt, Germany, 2000. (32) X-SHAPE 1.06: Crystal Optimization for Numerical Absorption Correction; STOE & Cie GmbH: Darmstadt, Germany, 1999. (33) Sheldrick, G. M. SHELXS-97: Program for Crystal Structure Solution; Universität Göttingen: Göttingen, Germany, 1997. (34) Sheldrick, G. M. Acta Crystallogr. 2008, A64, 112−122. (35) Sheldrick, G. M. SHELXL-97: Program for Crystal Structure Refinement; Universität Göttingen: Göttingen, Germany, 1997. (36) Brandenburg, K. Diamond 3.2g; Crystal Impact GbR: Bonn, Germany, 1999−2011. (37) Kohn, W.; Sham, L. J. Phys. Rev. A 1965, 140, 1133−1138. (38) (a) Becke, A. D. Phys. Rev. A 1988, 38, 3098−3100. (b) Becke, A. D. J. Chem. Phys. 1993, 98, 5648−5652. (c) Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B 1988, 37, 785−789. (39) For the basis sets and the corresponding literature, see: http:// www.emsl.pnl.gov/forms/basisform.html.In. (40) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman,

ASSOCIATED CONTENT

S Supporting Information *

A synthetic procedure for Hg(C6F5)2; spectroscopic data for Hg(C6F5)2 and C6F5HgCl; tables containing the energies as well as a table with selected bond lengths calculated for the species discussed in the theoretical sections; figures of the (−)-ESI mass spectra, of the crystal structures of Cs21·2Et2O, Cs23·2MeCN, and Cs24·4Me2CO, of the calculated structures of the dianions 1−5, and of the 11B NMR spectra of anions 6− 10; and CIF files giving details of the X-ray crystallographic analyses and the crystal data. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author *E-mail: [email protected] (J.W.), [email protected] (M.F.).



ACKNOWLEDGMENTS This work was supported by the Deutsche Forschungsgemeinschaft (FI 1628/2-1). The authors thank Professor W. Frank for generous support and Dr. G. Reiss, Mrs. E. Hammes and Mr. P. Roloff as well as Dr. P. Tommes (all Heinrich-HeineUniversität Düsseldorf), Dr. R. Bertermann (Julius-Maximilians-Universität Würzburg), Dr. Thomas Dülcks, and Dorit Kemken (both Universität Bremen) for technical support and helpful discussions.



REFERENCES

(1) (a) Bresadola, S. In Metal Interactions with Boron Clusters; Grimes, R. N., Ed.; Plenum Press: New York, 1982; pp 173−237. (b) Bregadze, V. I. Chem. Rev. 1992, 92, 209−223. (2) (a) Hawthorne, M. F.; Zheng, Z. P. Acc. Chem. Res. 1997, 30, 267−276. (b) Wedge, T. J.; Hawthorne, M. F. Coord. Chem. Rev. 2003, 240, 111−128. (3) (a) Lee, H.; Knobler, C. B.; Hawthorne, M. F. Angew. Chem. 2001, 113, 3148−3150;(b) Angew. Chem., Int. Ed. 2001, 40, 3058− 3060. (c) Lee, H.; Knobler, C. B.; Hawthorne, M. F. J. Am. Chem. Soc. 2001, 123, 8543−8549. (d) Lee, H.; Diaz, M.; Knobler, C. B.; Hawthorne, M. F. Angew. Chem. 2000, 112, 792−794;(e) Angew. Chem., Int. Ed. 2000, 39, 776−778. (f) Zheng, Z. P.; Knobler, C. B.; Hawthorne, M. F. J. Am. Chem. Soc. 1995, 117, 5105−5113. (4) Schöberl, U.; Magnera, T. F.; Harrison, R. M.; Fleischer, F.; Pflug, J. L.; Schwab, P. F. H.; Meng, X. S.; Lipiak, D.; Noll, B. C.; Allured, V. S.; Rudalevige, T.; Lee, S.; Michl, J. J. Am. Chem. Soc. 1997, 119, 3907−3917. (5) Taylor, T. J.; Burress, C. N.; Gabbaï, F. P. Organometallics 2007, 26, 5252−5263. (6) Shur, V. B.; Tikhonova, I. A. Russ. Chem. Bull. 2003, 52, 2539− 2554. (7) Körbe, S.; Schreiber, P. J.; Michl, J. Chem. Rev. 2006, 106, 5208− 5249. (8) Ivanov, S. V.; Rockwell, J. J.; Polyakov, O. G.; Gaudinski, C. M.; Anderson, O. P.; Solntsev, K. A.; Strauss, S. H. J. Am. Chem. Soc. 1998, 120, 4224−4225. (9) Tsang, C.-W.; Yang, Q.; Mak, T. C. W.; Xie, Z. Appl. Organomet. Chem. 2003, 17, 449−452. (10) Finze, M.; Sprenger, J. A. P. Chem.Eur. J. 2009, 15, 9918− 9927. (11) Himmelspach, A.; Zähres, M.; Finze, M. Inorg. Chem. 2011, 50, 3186−3188. (12) (a) Finze, M. Angew. Chem. 2007, 119, 9036−9039;(b) Angew. Chem., Int. Ed. 2007, 46, 8880−8882. (c) Finze, M.; Sprenger, J. A. P. Z. Anorg. Allg. Chem. 2010, 636, 1538−1542. 1576

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577

Organometallics

Article

J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 03, revision D.01; Gaussian, Inc.: Wallingford, CT, 2004. (41) Dirac, P. A. M. Proc. R. Soc. London, Ser. A 1929, 123, 714−733. Slater, J. C. Phys. Rev. 1951, 81, 385−390. (42) Vosko, S.; Wilk, L.; Nusair, M. Can. J. Phys. 1980, 58, 1200− 1211. (43) Møller, C.; Plesset, M. S. Phys. Rev. 1934, 46, 618−622. (44) (a) Turbomole V6.0 2009; a development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989−2007, TURBOMOLE GmbH: Karlsruhe, Germany, since 2007. Available from http://www.turbomole.com. (b) Ahlrichs, R.; Bär, M.; Häser, M.; Horn, H.; Kölmel, C. Chem. Phys. Lett. 1989, 162, 165−169. (45) (a) Eichkorn, K.; Treutler, O.; Ö hm, H.; Häser, M.; Ahlrichs, R. Chem. Phys. Lett. 1995, 240, 283−290. (b) Sierka, M.; Hogekamp, A.; Ahlrichs, R. J. Chem. Phys. 2003, 118, 9136−9148. (46) Hättig, C.; Weigend, F. J. Chem. Phys. 2000, 113, 5154−5161. (47) The corresponding literature for the basis sets can be found at ftp://ftp.chemie.uni-karlsruhe.de/pub/basen.

1577

dx.doi.org/10.1021/om201023h | Organometallics 2012, 31, 1566−1577